You are on page 1of 12

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/277916169

Experimental verification of thermodynamic


fatigue life prediction model using entropy as
damage metric

Article in Materials Science and Technology · May 2015


DOI: 10.1179/1743284715Y.0000000074

CITATIONS READS

3 26

2 authors, including:

Cemal Basaran
University at Buffalo, The State University of New York
210 PUBLICATIONS 2,726 CITATIONS

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

A New Mechanics Theory with Past, Present and Future variables: Unified Theory of Mechanics and
Thermodynamics View project

All content following this page was uploaded by Cemal Basaran on 15 January 2016.

The user has requested enhancement of the downloaded file. All in-text references underlined in blue are added to the original document
and are linked to publications on ResearchGate, letting you access and read them immediately.
Experimental Verification of a Thermodynamic Fatigue Life Prediction Model
Therence Temfack1, Cemal Basaran2*
1
Department of Mechanical and Aerospace Engineering
2
Department of Civil, Structural and Environmental Engineering
State University of New York at Buffalo
Buffalo, NY, USA 14260
*
Corresponding author, 2cjb@buffalo.edu

Keywords: Damage Mechanics, Fatigue, Thermodynamics, Entropy, Lifetime Prediction.

Abstract. A fatigue model based on entropy generation is presented and validated through experiments.
This model is purely physical and combines statistical mechanics with thermodynamic laws applied at a
local scale. The model does not require an empirical damage surface or phenomenological constitutive
modeling constants. Damage evolution parameter varies from 0 to 1. As it is for the irreversible internal
entropy production, this parameter is a non-decreasing quantity that increases with the degradation of a
material.

Introduction
All components and structures subjected to cycling loading are prone to fatigue failure. Micro
crack formation, coalescence and growth weakens the specimen, and leads to final failure which is
referred as fatigue, especially when then specimen undergoes cyclic loading. Many attempts have been
made to define and predict the fatigue evolution of materials in different circumstances [2-8], the well-
known Coffin-Manson law is one them. Solely, based on experimental data, Coffin-Manson model relates
the amplitude of the applied cyclic plastic deformation to the number of cycles to failure. This model has
been proven to be reliable for low-cycle fatigue regime, and the universality of this law leads us to
consider that there might be a general physical process that governs the observed fatigue phenomenon
[7-40]. But the Coffin-Manson and similar phenomenological models are empirical and do not give a
clue about the physical process underlying fatigue mechanism.
Materials experience low-cycle fatigue when subjected to plastic deformation and undergo high
cycle fatigue when subjected to elastic deformation only. It is possible to explain fatigue based on
thermodynamic principles. If any component or structure is considered to be a thermodynamic system, it
cannot be 100% efficient, as such, it must have a limited life span. It is also possible to explain fatigue
at the atomic scale, while this is very easy and obvious in the low-cycle fatigue regime, it is more complex
in the high cycle fatigue regime. Because according to continuum mechanics rules in the elastic region
when the material is unloaded it does not experience any permanent deformation and stress strain diagram
goes back to the origin. However this is not true when we inspect the material at the atomic scale. In the
elastic region under cycling loading when atomic bonds are stretched after unloading atoms do not
necessarily go back to their initial lattice site but to a site that is at a lower energy level [28]. This energy
minimization process is responsible for initiating nano-scale voids in the material. Of course these voids
because stress concentration points and as a result nanoscale voids grow and coalesce to form microscale
voids (cracks). Overtime micro cracks grow and coalesce into bigger cracks and eventually load carrying
intact cross section is subjected to a stress beyond the ultimate stress of the material and it fails. In

1
summary it is possible to describe fatigue process as an energy minimization process in the material at
the atomic scale. Of course this process happens much faster in low-cycle fatigue regime when the
applied stress is beyond the yield stress of the material and dislocation generation process accelerates
micro crack growth.
In other words, we believe that it is possible to explain fatigue as result of thermodynamics
inefficiency of the system (it established thermodynamics fact that 100% efficient engine is not possible).
Especially in the case of cycling loading under elastic stresses, if all atoms go back to their initial site
after unloading, as assumed by the continuum mechanics, fatigue would not be possible under elastic
regime. However that would violate the Clausius-Duhem inequality and laws of thermodynamics.
The process related to fatigue must involve permanent degradation due to irreversible processes
that lead to failure. Naderi et al. [1] and Liakat et al.[27] performed several fatigue test that show that the
cumulative entropy generation of specimens is constant at the time of failure and independent of load,
geometry and frequency. The original idea of this work is the following: entropy being the amount of
energy lost and no more available in a system, one can assume that this thermodynamic quantity is related
to the fatigue evolution of the specimen [2 and 8].
In order to derive a physical fatigue evolution model, we will first present the laws of
thermodynamics which constitute the foundations of this theory. The internal entropy generation of the
system given by the thermodynamics laws will then be used to derive a damage evolution parameter that
can be considered as a lifetime predictor. The model is finally applied to two cases: uniaxial cyclic
loading and monotonic traction. For each case, the damage parameter is calculated using experimental
data obtained in the laboratory by the authors on steel samples.

Table 1 Definition of Variables Used

NOMENCLATURE
volume specific entropy
surface boundary entropy flux
density specific internal entropy production
specific internal heat source per unit time temperature
heat flux recoverable energy
total energy flux plastic strain rate tensor
specific total energy elastic strain rate tensor
specific internal energy strain rate tensor
specific potential energy internal specific entropy
velocity molar mass
stress tensor gas constant
body force per unit mass disorder function

Thermodynamic Laws
There are four laws of thermodynamics to describe how fundamental physical quantities behave
under various circumstances. These laws define the evolution of thermodynamic systems. For the purpose
of this work, we only consider the first two laws that essentially relate temperature, energy and entropy
of a system. These laws must be applied at a local scale to be combined with statistical mechanics [10]

2
Energy Conservation. The first law of thermodynamics expresses the change of the total energy of a
system as it follows:
V Ω V
d
∭ ρe dV = - ∬ ⃗⃗Je dΩ
⃗⃗ + ∭ ρr dV . (1)
dt

Where the definitions of variables are given in Table 1.


This law simply states that the variation of total energy of system is due to internal heat source r and
energy flux ⃗⃗Je across the surface boundary [2]. The negative sign of the energy flux is justified by the fact
that the normal to the surface are outwards. Therefore, a negative energy flux is a positive external input
of energy to the system. Taken at the local scale, this law becomes:

∂ρe
= - div ⃗⃗⃗
Je + ρr . (2)
∂t

The total energy is summation of internal energy, potential energy and kinetic energy and [2]:

⃗v 2
e=u+Ψ+ . (3)
2

These energies are not conserved but converted from one to another so that the total is conserved. As we
will see later, the entropy generation is only linked to the internal energy [2]. Hence, we then need to
derive the internal energy evolution. The variation of kinetic and potential energies is derived from the
local conservation of momentum [2]:

dv

ρ = - ⃗⃗⃗⃗⃗ ⃗
divσ + ρb (4)
dt

Assuming a conservative body force derived from the potential energy:

𝜕Ψ
⃗b = - ⃗⃗⃗⃗⃗⃗⃗⃗⃗
gradΨ , = 0. (5)
𝜕𝑡

The local conservation of momentum can be defined by the rate of change of kinetic energy and potential
energy:

2
∂ρ(Ψ+ v
⃗ ⁄ )
2 = - div {ρ (Ψ + ⃗v 2⁄ ) v
∂t 2 ⃗ - σv
⃗ } - σ:D . (6)

By plugging Eq. 3 into Eq. 6 we find a new expression for the energy conservation:

∂ρe ∂ρu 2
= - div {ρ (Ψ + ⃗v ⁄2) v
⃗ - σv
⃗ } - σ:D . (7)
∂t ∂t

Knowing the relation between heat flux (J⃗⃗q ) and total energy flux (J⃗⃗e ):

3
J⃗⃗e = ρev ⃗ + ⃗⃗Jq .
⃗ - σv (8)

The internal energy evolution is derived by setting Eq. 7 and Eq. 2 equal.

∂ρu
⃗ + ⃗⃗Jq } + σ:D + ρr .
= - div{ρuv (9)
∂t

Using the following mathematical relation where α is a scalar value:

dα ∂αρ
ρ = + div(αρv
⃗). (10)
dt ∂t

Eq. 9 gives us another expression of the internal energy evolution by replacing α by u:

du
ρ = - div ⃗⃗Jq + σ:D + ρr . (11)
dt

Clausius-Duheim Inequity. The second law of thermodynamics, also called Clausius-Duheim inequity
states that the internal entropy production is a positive quantity. Before writing that inequity, we will first
express the entropy production. As for energy, the change in total entropy of the system is due to internal
entropy source and total entropy flux across the surface boundary Ω:
d V Ω V
dt
∭ ρs dV = - ∬ J⃗⃗⃗⃗⃗⃗⃗ ⃗⃗
s,tot dΩ + ∭ γ dV . (12)

Taken at the local scale:

∂ρs
= - div ⃗⃗⃗⃗⃗⃗⃗
Js,tot + γ . (13)
∂t

By rewriting the total entropy flux ⃗⃗⃗⃗⃗⃗⃗


Js,tot as the sum of entropy flux ⃗Js and a convective term ρsv
⃗ and with the
help of Eq. 10, Eq. 13 can be transformed into:

ds
ρ = - div ⃗Js + γ . (14)
dt

The entropy represents the energy that is lost and no more available to the system to do work. The entropy
and internal energy are related by:

Tds = du - dωe . (15)

By using Eq. 15, Eq. 11 becomes:

ds 1 dωe
ρ = ( - div ⃗⃗Jq + σ:D + ρr - ). (16)
dt T dt

Eq. 14 and Eq. 16 both express the total entropy variation. By equalizing them we can find the internal
entropy generation. The relation between heat flux and entropy flux is
4
⃗⃗Jq
J⃗s = . (17)
T

Using Eq. 17, Eq. 14 becomes:

ds ⃗⃗Jq
ρ = - div ( ) + γ . (18)
dt T

By mean of the following relation where α can be any scalar and ⃗θ any vector:


θ div(θ⃗) θ⃗
div ( ) = ⃗⃗⃗⃗⃗⃗⃗⃗⃗ (α) .
- 2 .grad (18)
α α α

Eq. 18 turns out to give:

ds ⃗⃗Jq divJ⃗⃗q
ρ = 2 ⃗⃗⃗⃗⃗⃗⃗⃗⃗
gradT - +γ. (20)
dt T T

Eq. 16 and Eq. 18 give the internal entropy generation of the system:

σ:D ρ dωe ⃗⃗Jq ρr


γ= - - 2 ⃗⃗⃗⃗⃗⃗⃗⃗⃗
gradT + . (21)
T T dt T T

In this model, small deformations are considered, which leads to:

dε ρdωe
D= , ε = εe + εp, = σ:ε̇ e (22)
dt dt

⃗⃗⃗⃗⃗⃗⃗⃗⃗ T. The internal


Furthermore, the heat flux is proportional to the gradient of temperature [2]: ⃗⃗Jq =-kgrad
entropy generation is then:

σ:ε̇ p k 2 ρr
γ= ⃗⃗⃗⃗⃗⃗⃗⃗⃗ T| +
+ 2 |grad . (23)
T T T

The Clausius-Duheim inequity expresses the irreversibility of the processes and we postulate that
damages (degradation) undergone by material is irreversible, as such:

γ≥0.

σ:ε̇ p k 2 ρr
⃗⃗⃗⃗⃗⃗⃗⃗⃗ T| +
+ 2 |grad ≥0. (24)
T T T

So far, we have used the first law of thermodynamics that expresses the total energy conservation
of a system to derive the variation of internal energy. From the internal energy and entropy variation of
5
the system, we derived an expression for the internal entropy production of the system. The latter quantity
is said to be positive according to the second law of thermodynamics. The internal entropy generation is
of utmost importance for the thermodynamic fatigue model proposed in here.
Entropy production represents the effects of irreversible processes in the system. In fact, the first
term, the plastic work contains the effects of mechanical loading. The second terms takes in account the
thermal loading while the last term represents any other type of loading undergone by the specimen.

Damage Evolution Parameter


The internal entropy generation derived previously is a way to put the effects of irreversible
processes undergone by the specimen. Now we are going to normalize this quantity to find a damage
evolution parameter for mechanical fatigue process. This damage parameter will be equal to 0 at the
initial state of life of the specimen, it will progressively increase when the specimen is loaded to reach
the value of 1 when failure occurs. This damage parameter will be derived via statistical mechanics.
In 1898, Boltzmann [11] first used statistical mechanics to give a precise meaning of disorder in
gas and established a connection between disorder and internal entropy for the whole system. The internal
entropy is expressed as a function of W, the disorder function that represents the probability that the system
will exist in the state relative to all the possible states it could be in [29]:

R
si = ln(β * W) . (25)
ms

A scale factor β is added to ensure that the mathematical and physical aspect of Eq. 25 are coherent. In
fact, si , R and ms are positive quantities (∈ R+ ) and the disorder function is a probability
(0 ≤ W ≤ 1 → ln(W) ≤ 0). β is then used to enforce the two sides of the equation to have the same sign.
We notice that Eq. 25 is equivalent to:

1 ms
W= ex p (si ). (26)
β R

We define a damage evolution parameter D which is defined as the ratio of the change in disorder
parameter with respect to the current state disorder parameter, (analogous to the true strain in continuum
mechanics)

W - W0 ms
D= = 1 - ex p (- (si - si0 ) ) . (27)
W R

One can notice that the scale factor β has vanished. With si0 the initial internal entropy of the system, so
that the initial damage is null, and for an infinite entropy, equivalent to failure, the damage is 1. This
damage parameter can be seen as the probability that the specimen will fail at the considered time of its
life. Of course it is well known that the definition of failure is different for each specific engineering
application. In most cases the system or component is assumed to fail long before D reaches the value of
one (1). We notice that this damage parameter depends on the internal entropy production through the
difference.
t
γ
(si - si0 ) = ∆si = ∫ dt . (28)
t0 ρ

6
t t t
σ:ε̇ p k 2 r
∆si = ∫ dt + ∫ ⃗⃗⃗⃗⃗⃗⃗⃗⃗ T| dt + ∫ dt .
|grad (29)
2
t 0 ρT t 0 ρT t0 T

Verification of the Model


Now that the damage evolution parameter is defined, the model will be applied to two cases and
validated through comparison with experimental data. First a fully reversed uniaxial cyclic loading is
considered and secondly a monotonic traction. The experiments were performed in the same conditions
and on the similar specimen (Fig. 1) for the two cases.

Fig. 1. Specimen dimensions and properties

In order to evaluate the damage evolution parameter (which is the entropy production rate) in both
cases, only the effects of mechanical load were taken in account. Considering Eq. 23, the second term
which represents the thermal effect vanishes if we assume that heat generated during cycling loading is
insignificant and the temperature of the specimen remains constant. The last term is not considered
because the specimen is only subjected to mechanical loading, tension-compression. Furthermore,
because the load is uniaxial, the problem is reduced to one direction and the plastic work is the product
of applied stress and induced plastic strain. The internal entropy generation is then reduced to:
t
σ:ε̇ p
∆si = ∫ dt . (30)
t 0 ρT

For the numerical computation, Eq. 30 is simplified to:

(σ*∆εp )
∆si = ∑ . (31)
ρT

Tension-Compression Cyclic Loading

The experiment consists of a displacement controlled test conducted in a MTS material


characterization unit, which applies uniaxial tension and compression in repeated and fully reversed
cycles to a sample. The experimental data obtained from the digital output consist of forces (F) and
displacements (u). The displacement’s range is [-1;1 mm] which corresponds to a 4% maximum strain in
the sample. Rate of loading is 0.5 Hertz. Here the failure occurs after 80 cycles. Since we assumed small
deformations in the model, plastic are calculated as it follows:
σ
εp = ε - εe , εp = ε - E
. (32)

7
Fig. 2. Stress-Strain diagram of cyclic loading Fig. 3. Damage parameter evolution for cyclic loading

Fig.2 shows the entire engineering stress and strain diagram obtained from the experiment which is a
fully reversed uniaxial cyclic loading. Figure 3 presents the damage evolution parameter as a function of
number cycles. As expected, the initial damage is null, and the damage increases with the number of
cycles and damage parameter goes to 1 when the specimen breaks.

Monotonic Loading Test

For this case, the sample is identical to the one used for the cyclic loading. The only difference is
that this time it is continuously loaded until failure occurs. The loading is a uniaxial tension, the
experimental data are similar and treated in the same way as before. Figure.4 shows the stress-strain
diagram obtained from experimental results of monotonic tension test performed on specimen. Figure 5
presents the evolution of the damage parameter for the specimen. As expected, the initial damage is null,
and the damage increases with the number of cycles and tends to 1 when the specimen breaks. But here,
we can see that the damage parameter remains very small for a while before it starts to increase. This
domain corresponds to the elastic range of the material. In fact, because we assumed small strain and
considered in the model that only plastic deformation could cause permanent degradation in the material,
there is no entropy generated in the elastic range according to Equation 24. Therefore, this model cannot
be used in case of high cycle fatigue. Because in that case, failure occurs under elastic deformation only.

8
Fig. 4. Stress-Strain diagram of cyclic loading Fig. 5. Damage parameter evolution for cyclic loading

Conclusions
A fatigue model using entropy generation as a damage metric has been proposed and validated
by cyclic loading and monotonic loading. The internal entropy generation was calculated through the
laws of thermodynamics and statistical mechanics. A component of a structure or even a whole structure
can be considered a thermodynamic system. The damage evolution parameter D proposed represents the
probability of failure of the considered component.
The model is adaptable to any type of loading undergone by the specimen (mechanical, thermal,
electrical, chemical …). It was shown that the model is suitable for lifetime of components subjected to
monotonic or cyclic loading. However, some assumptions that are valid most of the time were made to
simplify the model. The model might need some improvement for some applications. Especially, it is
assumed that elastic deformation cannot cause failure, which is not true. But compared to those of plastic
deformation, the effects of elastic deformation are negligible and can be assumed to be null.
For practical use, the damage evolution parameter derived here must be multiplied by a safety
coefficient. In fact, components must be checked and replaced long before their failure. The safety
coefficient may vary with the application. For instance, aircraft design and bicycle design do not require
the same level of factor of safety constraints.

References
1. M. Naderi, M. Amiri, M.M. Khonsari, On the thermodynamic entropy of fatigue fracture,
Proceedings of the Royal Society A: Mathematical, Physical and Engineering Science, 466 (2009)
423–438.
2. C. Basaran and S. Nie, An Irreversible Thermodynamic Theory for Damage Mechanics of Solids,
International Journal of Damage Mechanics, 13 (2004) 205-224.
3. C. Basaran, S. Nie, J. Gomez, E. Gunel, S. Li, M. Lin, H. Tang, C. Yan, W. Yao, H. Ye,
Thermodynamic theory for damage evolution in solids, in Handbook of Damage Mechanics,
2014, pp 721-762.
4. H. Tang and C. Basaran, A Damage Mechanics Based Fatigue Life Prediction Model for solder
joints, Trans. Of ASME, Journal of Electronic Packaging, 125 (2003) 120-125.

9
5. D. Kim, G. Dargush and C. Basaran, A cyclic two-surface thermoplastic damage model with
application to metallic plate dampers, Engineering Structures, 52 (2013) 608–620.
6. C. Basaran and S. Nie, A Thermodynamics Based Damage Mechanics Model for Particulate
Composites, International Journal of Solids and Structures, 44 (2007) 1099-1114.
7. M. Amiri and M.M. Khonsari, On the role of entropy generation in processes involving fatigue,
Entropy, 14 (2012) 24-31.
8. Basaran, C., and Yan. C.Y."A Thermodynamic Framework for Damage Mechanics of Solder
Joints,” Trans. of ASME, Journal of Electronic Packaging, Vol. 120, 379-384, 1998.
9. Basaran, C. and Tang, H.,”Implementation of a Thermodynamic Framework for Damage
Mechanics of Solder Interconnects in Microelectronic Packaging,” International Journal of
Damage Mechanics, Vol. 11, No. 1, pp. 87-108, January 2002.
10. Ye, H., Basaran, C. and Hopkins, D., Damage Mechanics of Microelectronics Solder Joints Under
High Current Densities,” International Journal of Solids and Structures, Vol. 40, No. 15, pp.
4021-4032, July 2003.
11. Basaran, C., Lin, M. and Ye. H. “A Thermodynamic Model for Electrical Current Induced
Damage”, Int. Journal of Solids and Structures Vol. 40, No 26 pp. 7315-7327, November 2003
12. Ye, H., Basaran, C. and Hopkins, D.,”Mechanical Degradation of Microelectronics Solder Joints
Under Current Stressing,” Int. Journal of Solids and Structures, Vol 40 No 26, pp 7269-7284,
November 2003
13. Gomez, J. and Basaran, C.”A Thermodynamics Based Damage Mechanics Constitutive Model
for Low Cycle Fatigue Analysis of Microelectronics Solder Joints Incorporating Size Effect,”
International Journal of Solids and Structures, Vol. 42, issue 13, pp. 3744-3772, (2005)
14. Basaran, C., Zhao, Y., Tang, H. and Gomez, J..”A Damage Mechanics Based Unified Constitutive
Model for Solder Alloys,” Trans of ASME, Journal of Electronic Packaging, Vol. 127, issue 3,
pp. 208-214, September 2005 .
15. Gomez, J. and Basaran, C., ”Damage Mechanics Constitutive Model for Pb/Sn Solder Joints
Incorporating Nonlinear Kinematic Hardening and Rate Dependent Effects Using a Return
Mapping Integration Algorithm,” Mechanics of Materials, 38 (2006), 585-598.
16. Gomez, J, Lin, M. and Basaran, C. “Damage Mechanics Modeling of Concurrent Thermal And
Vibration Loading On Electronics Packaging” Multidiscipline Modeling in Materials and
Structures, Vol.2, No.3, pp. 309-326, July 2006.
17. Basaran, C. and Nie, S.“A Thermodynamics Based Damage Mechanics Model for Particulate
Composites,” International Journal of Solids and Structures, 44, (2007) 1099-1114.
18. Basaran, C. and Lin, M.,” Damage Mechanics of Electromigration in Microelectronics Copper
Interconnects,” International Journal of Materials and Structural Integrity, vol. 1, Nos1/2/3, 16-
39, 2007.
19. Basaran, C. and Lin, M.,” Damage Mechanics of Electromigration Induced Failure,” Mechanics
of Materials, 40 (2008) 66-79,
20. Li, S., Abdulhamid, M. and Basaran, C. “Simulating Damage Mechanics of Electromigration
and Thermomigration,” Simulation: Transactions of the Society for Modeling and Simulation
International Vol. 84, Number 8/9, pp. 391-401, August/September 2008.

10
21. Gomez, J. and Basaran, C. “Computational Implementation of Cosserat Continuum,”,
International Journal of Materials and Production Technology, Vol. 34, No 1 / 2, pp 3-36,
January 2009,
22. Li, S. and Basaran, C. “A Computational Damage Mechanics Model for Thermomigration,”
Mechanics of Materials, vol. 41, issue 3, pp. 271-278, March 2009,
23. Li, S., Abdulhamid, M., Basaran, C. and Lai, Y.S. “Damage Mechanics of Low Temperature
Electromigration and Thermomigration” IEEE Trans. On Advanced Packaging, Vol. 32, No 2,
pp., 478-485, May 2009
24. Li, S., Sellers, Basaran, C., M. Schultz, A.,. and Kofke, D., ”Lattice Strain Due to an Atomic
Vacancy,” Int. J. of Molecular Sciences, 2009, 10, 2798-2808,
25. Yao, W. and Basaran, C., ” Electromigration Analysis of Solder Joints under AC Load: a Mean
Time to Failure Model” J. of Applied Physics, 19 March 2012, Journal of Applied Physics ,
Vol.111, Issue 6,
26. M. Amiri and M. Modarres, An entropy-based damage characterization, Entropy, 16 (2014) 6434-
6463.
27. M. Liakat and M.M. Khonsari, Entropic characterization of metal fatigue with stress
concentration, International Journal of Fatigue, 70 (2015) 223-234.
28. Seller, M., Schultz, A., Basaran, C. and Kofke, D., “Sn Grain Boundary Structure and Self-
Diffusivity via Molecular Dynamics Simulation” Physical Review B, 81, 134111
29. Boltzmann, I. Lectures on gas theory, University of California Press, Berkeley, CA, 1898.

11

You might also like